Skip to main content
  • Research article
  • Open access
  • Published:

HDAC6 regulates dental mesenchymal stem cells and osteoclast differentiation

Abstract

Background

Dental and periodontal tissue development is a complicated process involving a finely regulated network of communication among various cell types. Understanding the mechanisms involved in regulating dental mesenchymal stem cells (MSCs) and osteoclast cell differentiation is critical. However, it is still unclear whether histone deacetylase HDAC6 is involved in dental MSCs fate determination and osteoclast differentiation.

Methods

We used shRNA and siRNA knockdown to explore the role of HDAC6 in dental MSCs odontogenic differentiation and osteoclasts maturation.

Results

Based on HDAC6 knockdown dental MSCs, our data suggest that HDAC6 knockdown significantly increases alkaline phosphate activity and mineralized nodules formation. Additionally, mRNA expression of odontogenic marker genes (OSX, OCN, and OPN) was induced by HDAC6 knockdown. By using HDAC6 siRNA, we knocked down HDAC6 in osteoclast precursor RAW 264.7 cells. Our data suggests that HDAC6 knockdown significantly inhibited osteoclasts differentiation. Additionally, mRNA expression of osteoclast marker genes Trap, Mmp9, and Ctsk was decreased by HDAC6 knockdown.

Conclusions

Our study demonstrated that HDAC6 plays an important role in regulating dental MSCs and osteoclasts differentiation.

Peer Review reports

Background

Bone is continuously undergoing resorption and remodeling to maintain bone volume and calcium homeostasis. Osteoblasts and osteoclasts play critical roles in regulating bone formation and resorption. Osteoblast cells are believed to derive from MSCs that originate in bone marrow [1]. MSCs are adult stem cells which can differentiate into chondrocytes, adipocytes, osteoblasts, myocytes, and hepatocytes [2, 3]. Osteoclasts are multinucleated cells differentiated from hematopoietic stem cells (HSCs) [4, 5]. Because the regulation of osteoclastogenesis and osteoblastogenesis are critical in tooth eruption and movements [6], many studies have been conducted to unveil the mechanism involved in bone cell differentiation. Broad-acting histone deacetylase (HDAC) inhibitor TAS has been reported to increase osteoblast proliferation and the transcriptional activity of Runx2 [7]. Another study demonstrated that HDAC inhibitor FR901228 decreased the maturation of osteoclasts [8]. However, there is still much to be understood about the specific role of individual HDACs in the regulation of dental bone resorption and regeneration.

Studies have shown that dental MSCs have multipotency for odontogenic, osteogenic, chondrogenic, and adipogenic potential under different culture conditions [9,10,11]. Therefore, dental MSCs are a promising treatment approach for clinical applications in dental regeneration and craniofacial therapies. The features of dental MSCs are influenced by various factors including cytokines, cell passages, and oxygen concentration. In the last decade, our knowledge about molecular contributors to MSCs differentiation has also increased. Bone lineage commitment is reportedly regulated by the expression of transcription factor RUNX2, which further promotes expression of alkaline phosphatase (ALP), osterix (OSX), osteopontin (OPN), and osteocalcin (OCN) [12]. In addition, several transcription factors and signal pathways have been demonstrated to regulate MSCs differentiation [13, 14].

Osteoclasts are multinucleated cells differentiated from monocytes or macrophages that absorb bone matrix. Osteoclast differentiation is regulated by the tumor necrosis factor (TNF) family cytokine, receptor activator of nuclear factor (NF)-κB ligand (RANKL), and macrophage colony-stimulating factor (M-CSF) [15,16,17]. Studies have suggested that TNF receptor-associated factor 6 (TRAF-6), NF-κB and c-fos are also essential molecules for osteoclast genesis [17]. In addition, genome-wide screening has provided insights into additional genes that are involved in the regulation of osteoclast differentiation.

Histone deacetylases (HDACs) are a group of conserved enzymes that remove acetyl groups from lysine side chains of both histones and nonhistone proteins. HDACs are reportedly involved in the regulation of bone formation and maintenance [18, 19]. HDAC6, a unique protein that belongs to class IIb of HDACs, can shuttle between the cytoplasm and nucleus. Previous studies have demonstrated that HDAC6 inhibitors accelerate osteoblast maturation and suppress osteoclast differentiation [20]. HDAC6-deficient mice develop a slightly enlarged tibia and increased bone mineral density [21]. The Rho–mDia2–HDAC6 pathway is reportedly involved with osteoclast maturation by controlling podosome patterning through microtubule acetylation in osteoclasts [22]. While these studies indicate a role for HDAC6 in bone growth and resorption, the role of HDAC6 in dental MSCs and osteoclast differentiation is still unclear. In this study, we demonstrate that HDAC6 knockdown in dental MSCs can promote osteoblast maturation. We further demonstrate that HDAC6 loss can inhibit osteoclast differentiation. Thus, our findings suggest that HDAC6 plays an important role in dental bone growth and maintenance.

Methods

Human DPSCs isolation and culture

All patients provided informed consent according to the guidelines of the Ethics Committee of Chinese PLA General Hospital. Dental pulp was obtained from normal third molars of three adults donors (19–29 years of age) at the Dental Clinic of Chinese PLA General Hospital. Human DPSCs were isolated as previously reported [23]. Cells were cultured in a humidified 5% CO2 incubator at 37 °C in alpha-modified Eagle’s medium (Invitrogen) supplemented with 15% fetal bovine serum (Invitrogen), Gibco MEM non-essential amino acids (Invitrogen), 2 mmol/L L-glutamate, 100 units/mL penicillin, and 100 units/mL streptomycin. The RAW 264.7 cell line was cultured in a humidified 5% CO2 incubator at 37 °C in RPMI-1640 medium (Invitrogen) supplemented with 5% fetal bovine serum, 100 units/ml penicillin, and 100 units/ml streptomycin.

Viral infection

Lentiviral expression vector was constructed using pLKO.1 vector (Addgene). The sequence for scramble (control) was 5’-CCTAAGGTTAAGTCGCCCTCG-3′. The target sequences for shRNA were HDAC6sh1, 5’-CATCCCATCCTGAATATCCTT-3′ and HDAC6sh2, 5’-GCACAGTCTTATGGATGGCTA-3′. The insert was subcloned into pLKO.1 using Agel and EcoRI sites. Lentivirus production was performed according to the protocol provided by Addgene. Forty-eight hours after the transfection, the media containing viruses were collected and concentrated by ultracentrifugation.

Dental MSCs were seeded overnight and then infected with lentivirus in the presence of polyybrene (6 μg/mL; Sigma-Aldrich) for 24 h. The cells were then selected with puromycin for 72 h. Puromycin-resistant clones were cultured, and HDAC6 expression was detected by real-time RT-PCR. In rescue experiments, HDAC6-knockdown dental MSCs were infected with adenovirus (Sigma-Aldrich, St. Louis, MO, USA) containing flag-tagged HDAC6.

siRNA transfection

siRNAs were synthesized by Invitrogen and transfected with a Lipofectamine 2000 reagent (Invitrogen) at a final concentration of 40 nM. The target sequences for siRNA were HDAC6siRNA1: 5’-GCACCAUGGUCAAGGAACA-3′ and HDAC6siRNA2: 5’-CCAAUCUAGCGGAGGUAAA -3′.

ALP staining and ALP activity assay

Dental MSCs were cultured in a osteogenic medium containing 100 μmol/L ascorbic acid, 2 mmol/L b-glycerophosphate, and 10 nmol/L dexamethasone. After induction, cells were washed with PBS and fixed with 4% paraformaldehyde and incubated in BCIP/NBT solution (Beyotime, Shanghai, China) in the dark. Areas that stained purple were regarded as positive. ALP activities were determined using p-nitropheyl phosphate (Sigma) as described previously [24].

Alizarin red staining (ARS)

Dental MSCs were cultured in osteogenic medium. After 3 weeks, cells were washed with PBS, and fixed with 70% ethanol at 4 for 1 h. Thereafter, the cells were stained with 2% Alizarin Red (Sigma-Aldrich). To quantify the calcium mineral deposition, Alizarin Red was destained with 10% cetylpyridinium chloride in 10 mmol/L sodium phosphate for 30 min at room temperature. Alizarin Red concentration was determined by absorbance at 562 nm on a microplate reader using a standard calcium curve in the same solution. The final calcium level in each group was normalized with the total protein concentration obtained from a duplicate plate.

Osteoclast induction and trap staining

RAW 264.7 Cells were seeded in a 12-well culture plate (Corning) with OC differentiation medium containing 100 ng/mL recombinant RANKL (PeproTech). After 6 days, the medium was removed and the cells were washed with PBS. The cells were then subjected to TRAP staining (Sigma-Aldrich) following the manufacturer’s instructions to confirm their OC identity.

Real-time RT-PCR

Total RNA was extracted using TRIzol reagents (Invitrogen) and was transcribed with PrimeScript RT reagent kit (Takara Bio, Inc., Kusatsu, Japan). cDNA amplification and detection were performed using the Bio-Rad iQ5 real-time PCR system (Bio-Rad, Hercules, CA, USA) using SYBR Premix Ex Taq kit (Takara) and specific primers. The primers sequences are listed below:

  • 18S rRNA-forward: 5’-CGGCTACCACATCCAAGGAA-3′;

  • 18S rRNA-reverse, 5’-GCTGGAATTACCGCGGCT-3′.

  • HDAC6-forward: 5’-TCAGGTCTACTGTGGTCGTT-3′;

  • HDAC6-reverse: 5’-TCTTCACATCTAGGAGAGCC-3′.

  • OSX-forward: 5’-CGCTTTGTGCCTTTGAAA-3′;

  • OSX-reverse: 5’-CCGTCAACGACGTTATGC-3′.

  • OCN-forward: 5’-CAGACACCATGAGGACCATC-3′;

  • OCN-reverse: 5’-GGACTGAGGCTCTGTGAGT-3′.

  • OPN-forward: 5’-ATGATGGCCGAGGTGATAGT-3′;

  • OPN-reverse: ACCATTCAACTCCTCGCTTT-3′.

  • ALP-forward: 5’-GACCTCCTCGGAAGACACTC-3′;

  • ALP-reverse: 5’-TGAAGGGCTTCTTGTCTGTG-3′.

  • Mouse β-actin-forward: GATGCCAGCGACAAGAGGTT-3′;

  • Mouse β-actin-reverse: 5’-CATACCAGGGGATGTTGCGAA-3′.

  • Mouse HDAC6-forward: 5’-GAAGGAGGAGCTGATGTTGG-3′;

  • Mouse HDAC6-reverse: 5’-TCATGTACTGGGTTGTCTCCAT-3′.

  • Mouse TRAP-forward: 5’-GAAGAAGACTCACCAGAAGCAG-3′;

  • Mouse TRAP-reverse: 5’-TCCAGGTTATGGGCAGAGATT-3′.

  • Mouse Mmp9-forward: 5’-CAGGAGAGGCATTATGAGCA-3′;

  • Mouse Mmp9-reverse: 5’-GGTACTTTCCTGGTTCGCAT-3′.

  • Mouse Ctsk-forward: 5’-CTGGACAGCCAGACACTAAAG-3′;

  • Mouse Ctsk-reverse: 5’-CTCGCGGCAAGTCTTCAGAG-3′.

Western blot

Cells were solubilized in ice-cold RIPA buffer (Invitrogen) supplemented with complete protease inhibitor tablets (Roche, Basel, Switzerland). 40 μg of total protein was resolved by SDS-PAGE and transferred to PVDF membranes (Millipore, Burlington, MA, USA). Membranes were blotted with antibodies against HDAC6 (Abcam, Cambridge, UK) and α-tubulin (Cell Signaling Technology, Danvers, MA, USA) and HRP-conjugated secondary antibodies. The protein bands were visualized by the ECL system according to the manufacturer’s instructions and exposed to x-ray film.

Statistics

Data are presented as the mean ± SD of at least three independent experiments. Differences between groups were evaluated using Student’s t-test. p-value < 0.05 was considered significant.

Results

HDAC6 inhibits odontogenic differentiation of dental MSCs.

Previous studies have suggested that the inhibition of HDAC6 could promote BMSCs differentiation. Dental MSCs are reportedly capable of differentiating into odontoblasts similar to BMSCs. Therefore, we designed experiments to explore the function of HDAC6 in dental MSCs differentiation. Dental MSCs were infected with lentivirus-carrying HDAC6 shRNAs to generate HDAC6 knockdown cell lines. The knockdown efficiency of two different HDAC6 shRNAs was confirmed by Western blot (Fig. 1a). As HDAC6 sh2 showed higher knockdown efficiency, we chose dental MSCs infected with HDAC sh2 for use in further experiments. When these cells were induced with bone formation medium, MSC/sh2 cells showed increased bone formation as demonstrated by ALP staining and activity on day 7 (Fig. 1b). Furthermore, HDAC6 knockdown also increased the formation of mineralized nodules after prolonged induction with bone formation medium for 21 days as demonstrated by ARS staining (Fig. 1c). By real-time RT-PCR, we examined mRNA expression of odontogenic marker genes at different time points after bone formation induction. Our results showed that HDAC6 knockdown significantly induced the expression of OSX (Fig. 1e), ALP (Fig. 1f), OCN (Fig. 1g), and OPN (Fig. 1h). Taken together, these results indicate that HDAC6 plays an important role in odontogenic differentiation of MSCs.

Fig. 1
figure 1

HDAC6 knockdown increases odontogenic potential of dental MSCs. a ShRNA knockdown efficiency of HDAC6 was confirmed by Western blot. b Knockdown of HDAC6 increased ALP activity in dental MSCs. c Knockdown of HDAC6 promoted mineralization in MSCs. d Knockdown of HDAC6 promoted OSX expression in dental MSCs. e Knockdown of HDAC6 promoted ALP expression in dental MSCs. f Knockdown of HDAC6 promoted OCN expression in dental MSCs. g Knockdown of HDAC6 promoted OPN expression in dental MSCs

Rescue of HDAC6 inhibits odontogenic differentiation of dental MSCs.

To validate our finding from HDAC6 knockdown dental MSCs, we restored HDAC6 expression in MSCs expressing HDAC6 shRNA by infecting the cells with lentivirus-expressing Flag-HDAC6. Expression of HDAC6 was confirmed by Western blot (Fig. 2a). When HDAC6 sh2/Flag-HDAC6 cells were cultured with osteogenic medium, the cells showed similar bone formation capacity to control cells (scrsh/V), whereas HDAC6 sh2/V cells showed impaired bone formation capacity in ALP (Fig. 2b) and ARS assays (Fig. 2c). Similarly, the expression of OSX, ALP, OCN, and OPN as a result of HDAC6 knockdown was impaired after HDAC6 expression was restored (Fig. 2d–g), indicating that HDAC6 inhibits odontogenic differentiation.

Fig. 2
figure 2

The rescue of HDAC6 inhibits odontogenic potential of dental MSCs. a Overexpression of HDAC6 in MSCs that express HDAC6 shRNA is confirmed by Western blot. b Overexpression of Flag-HDAC6 inhibited ALP activities in MSC differentiation. c Overexpression of Flag-HDAC6 inhibited mineralization of MSC differentiation. d Overexpression of Flag-HDAC6 decreased OSX expression in MSC differentiation. e Overexpression of Flag-HDAC6 decreased ALP expression in MSC differentiation. f Overexpression of Flag-HDAC6 decreased OCN expression in MSC differentiation. g Overexpression of Flag-HDAC6 decreased OPN expression in MSC differentiation

Knockdown of HDAC6 inhibits osteoclast cells differentiation

HDAC6 inhibitors inhibit osteoclast cell differentiation. Therefore, we designed an experiment to study the role of HDAC6 in osteoclast differentiation. Macrophage line were transfected with control siRNA and HDAC6 siRNAs. The knockdown efficiency of HDAC6 was detected by Western blot (Fig. 3a). We chose HDAC6 siRNA2 with higher knockdown efficiency for further experiments. After transfection with HDAC6 siRNAs, RAW 264.7 cells were induced with RANKL to activate differentiation. Compared with the control line, RAW 264.7 cells transfected with HDAC6 siRNA showed a lower capability to form mature osteoclasts (Fig. 3b). Interestingly, the expression of HDAC6 was also induced by RANKL in RAW 264.7 cells transfected with control siRNA (Fig. 3c). Trap, Mmp9, and Ctsk genes play critical roles in osteoclast differentiation [25,26,27] Furthermore, we found that RAW 264.7 cells transfected with HDAC6 siRNA2 showed lower expression levels of Trap (Fig. 3d), Mmp9 (Fig. 3e), and Ctsk (Fig. 3f).

Fig. 3
figure 3

HDAC6 knockdown inhibited osteoclasts differentiation. a siRNA knockdown of HDAC6 is confirmed by Western blot. b Knockdown of HDAC6 inhibited osteoclast differentiation. c Expression of HDAC6 was induced by RANKL in RAW 264.7 cells. d Knockdown of HDAC6 decreased Trap expression. e Knockdown of HDAC6 decreased Mmp9 expression. f Knockdown of HDAC6 decreased Ctsk expression

Rescue of HDAC6 restores osteoclast differentiation

We restored HDAC6 expression in RAW 264.7 cells transfected with HDAC6 siRNA by infecting the cells with adenoviruses expressing Flag-HDAC6. HDAC6 expression was confirmed by Western blot (Fig. 4a). When HDAC6 siRNA/Flag-HDAC6 cells were induced with RANKL, the cells showed similar osteoclast differentiation capacity to the control cells (control siRNA/V) in a trap staining assay (Fig. 4b). Similarly, the decreased expression of Trap, Mmp9, and Ctsk as a result of HDAC6 knockdown was rescued after HDAC6 expression was restored (Fig. 4c–e), indicating that HDAC6 promotes osteoclast differentiation.

Fig. 4
figure 4

Rescue of HDAC6 restored osteoclast differentiation. a Overexpression of HDAC6 in osteoclasts that transfected with HDAC6 siRNA. b Overexpression of HDAC6 restored osteoclast differentiation. c Overexpression of HDAC6 increased Trap expression. d Overexpression of HDAC6 increased Mmp9 expression. e Overexpression of HDAC6 increased Ctsk expression

Discussion

Using dental MSCs to regenerate forms and functions of teeth is one of the important goals in dental tissue engineering, although additional fields of application are quickly emerging. Previous studies have discovered mechanisms involved in regulating odontogenic differentiation potential in dental MSCs [28, 29]. While dental MSCs are similar to BMSCs in their potential to differentiate into mineralized tissues, dental MSCs have more restricted differentiation than bone marrow cells have in vivo. It has been reported that dental MSCs are more committed to odontogenic than to osteogenic development in vitro [30]. Stem cells derived from dental tissue have been used for tissue engineering studies to assess their potential in pre-clinical application [31,32,33,34]. Therefore, a better understanding of the molecular mechanisms underlying cell differentiation of dental MSCs is of significant interest in regenerative medicine.

Osteoclasts are highly specialized migratory cells that regulate bone resorption. They are critical for normal skeletal growth and the maintenance of bone integrity throughout life. Overactive osteoclasts may cause several bone diseases including Paget’s disease in bone, juvenile Paget’s disease, expansile skeletal hyperphosphatasia, and familial expansile osteolysis [35,36,37,38,39]. Dental problems resulting from osteoclast activation have long been recognized. To improve our understanding and treatment of dental abnormalities and osteoclast diseases, it is critical to study mechanisms involved in regulation of osteoclasts differentiation.

HDACs can induce specific changes in gene expression by deacetylating both histone and non-histone proteins. HDAC inhibitors have been used in clinics for treatment of cancer, epilepsy, and bipolar disorder. In 1993, Iwami and Moriyama demonstrated that the HDAC inhibitor NaB could induce ALP expression in a MC3T3-E1 pre-osteoblast cell line [40]. Several studies suggest the HDAC inhibitor TSA has stimulatory effects on osteoblast cell lines, primary calvarial osteoblasts, and in calvarial organ cultures [7, 41]. TSA and NaB inhibit RANKL-mediated osteoclast differentiation from hematopoietic precursors [42]. Furthermore, several HDACs play important roles in bone development and physiology; HDAC1, HDAC3, HDAC6, and HDAC7 are involved with osteoblastogenesis through cooperation with, or inhibition of, Runx2 [43,44,45,46]. There is still much to learn about the roles of HDACs in osteoclast differentiation. Interestingly, a study from 2011 showed that HDAC3 and HDAC7 play opposite roles in osteoclast differentiation [47]. Suppression of HDAC7 facilitated osteoclastogenesis and increased osteoclast size [48]. HDAC9 knockout mice had elevated osteoclast numbers and bone resorption indices [49]. Class III HDAC Sirt1 also inhibited osteoclast differentiation in a conditional deletion Sirt1 mouse model [50].

As a member of the class II HDAC family, HDAC6 predominantly localizes in the cytoplasm. It plays an important role in regulating cell motility by deacetylating a-tubulin and cortactin [51, 52]. HDAC6 also modulates Hsp90-dependent activation of the glucocorticoid receptor through deacetylating Hsp90 [53, 54]. Loss of HDAC6 increased the formation of ossified bone in TDII embryos, but it did not lead to significantly different bone mineral densities at later ages [55, 56]. Specific inhibition of HDAC6 with Tubastatin or AC1215 increased osteoblast formation and osteoclast inhibition [20]. Taking into account previous published results, we have shown that HDAC6 knockdown promotes dental MSCs differentiation by using shRNA. By employing a small interfering RNA technique, we also knocked down the expression of HDAC6 in murine osteoclast precursor cells line RAW 264.7. The knockdown of HDAC6 in RAW 264.7 cells inhibited osteoclasts formation under RANKL stimulation. However, it is still not clear whether HDAC6 could regulate osteoclast precursor proliferation. Although further studies are required to understand the detailed molecular mechanisms by which HDAC6 regulates the expression of osteoclast marker genes, our results shed light on the role of HDAC6 in dental MSCs and osteoclasts differentiation. Together with previous findings, our study suggests that HDAC6 is a potential regulator for bone growth and maintenance.

Conclusions

We expect that HDAC6 knockdown will lead to inhibition of osteoclastogenesis and the promotion of dental MSCs odontogenic differentiation. Our findings give new insight into the role of HDACs in tooth development and disease and suggest HDAC6 as a novel therapeutic target for disease treatment and drug development. HDAC6 knockout mice appear to be healthy, and highly specific HDAC inhibitors appear to be well tolerated in clinical trials. The inhibition of HDAC6 may offer health advantages under some dental conditions. In future studies, it will be important to determine how HDAC6 is involved with networks regulating dental cell differentiation.

Abbreviations

ALP:

Alkaline phosphatase

ARS:

Alizarin red staining

HDACs:

Histone deacetylases

MSCs:

Mesenchymal stem cells

OCN:

Osteocalcin

OPN:

Osteopontin

OSX:

Osterix

RANKL:

Receptor activator of nuclear factor (NF)-κB Ligand

TNF:

Tumor necrosis factor

TRAF-6:

Receptor-associated factor 6

References

  1. Undale AH, Westendorf JJ, Yaszemski MJ, Khosla S. Mesenchymal stem cells for bone repair and metabolic bone diseases. Mayo Clin Proc. 2009;84:893–902.

    Article  Google Scholar 

  2. Hoffmuller U. MSC 2007--Adult Mesenchymal Stem Cells in Regenerative Medicine. Developments in stem cell therapeutic research. IDrugs. 2007;10(11):787–90.

  3. Kaiser S, Hackanson B, Follo M, Mehlhorn A, Geiger K, et al. BM cells giving rise to MSC in culture have a heterogeneous CD34 and CD45 phenotype. Cytotherapy. 2007;9:439–50.

    Article  Google Scholar 

  4. Kukita T, Kukita A, Iijima T. Osteoclast differentiation antigen involved the induction of calcitonin escape phenomenon. Kaibogaku Zasshi. 2000;75:433–8.

    PubMed  Google Scholar 

  5. Shalhoub V, Elliott G, Chiu L, Manoukian R, Kelley M, et al. Characterization of osteoclast precursors in human blood. Br J Haematol. 2000;111:501–12.

    Article  Google Scholar 

  6. Cantley MD, Zannettino ACW, Bartold PM, Fairlie DP, Haynes DR. Histone deacetylases (HDAC) in physiological and pathological bone remodelling. Bone. 2017;95:162–74.

    Article  Google Scholar 

  7. Schroeder TM, Westendorf JJ. Histone deacetylase inhibitors promote osteoblast maturation. J Bone Miner Res. 2005;20:2254–63.

    Article  Google Scholar 

  8. Nakamura T, Kukita T, Shobuike T, Nagata K, Wu Z, et al. Inhibition of histone deacetylase suppresses osteoclastogenesis and bone destruction by inducing IFN-beta production. J Immunol. 2005;175:5809–16.

    Article  Google Scholar 

  9. Holt RD, Oliver M. Evaluating web-based learning modules during an MSc programme in dental public health: a case study. Br Dent J. 2002;193:283–6.

    Article  Google Scholar 

  10. Calvert G, Britten N. The united medical and dental School of Guy's and St Thomas' Hospitals' MSc in general practice: graduates' perspectives. Med Educ. 1999;33:130–5.

    Article  Google Scholar 

  11. Smith GP. Formula for success: the role of the MSC officer and the U.S. Army dental care system. Mil Med. 1991;156:134–6.

    Article  Google Scholar 

  12. Lian JB, Javed A, Zaidi SK, Lengner C, Montecino M, et al. Regulatory controls for osteoblast growth and differentiation: role of Runx/Cbfa/AML factors. Crit Rev Eukaryot Gene Expr. 2004;14:1–41.

    Article  Google Scholar 

  13. Gharibi B, Ghuman MS, Hughes FJ. Akt- and Erk-mediated regulation proliferation and differentiation during PDGFRbeta-induced MSC self-renewal. J Cell Mol Med. 2012;16:2789–801.

    Article  Google Scholar 

  14. Roche S, Richard MJ, Favrot MC. Oct-4, rex-1, and Gata-4 expression in human MSC increase the differentiation efficiency but not hTERT expression. J Cell Biochem. 2007;101:271–80.

    Article  Google Scholar 

  15. Srivastava S, Toraldo G, Weitzmann MN, Cenci S, Ross FP, et al. Estrogen decreases osteoclast formation by down-regulating receptor activator of NF-kappa B ligand (RANKL)-induced JNK activation. J Biol Chem. 2001;276:8836–40.

    Article  Google Scholar 

  16. Neale SD, Smith R, Wass JA, Athanasou NA. Osteoclast differentiation from circulating mononuclear precursors in Paget's disease is hypersensitive to 1,25-dihydroxyvitamin D(3) and RANKL. Bone. 2000;27:409–16.

    Article  Google Scholar 

  17. Kobayashi K, Takahashi N, Jimi E, Udagawa N, Takami M, et al. Tumor necrosis factor alpha stimulates osteoclast differentiation by a mechanism independent of the ODF/RANKL-RANK interaction. J Exp Med. 2000;191:275–86.

    Article  Google Scholar 

  18. Karamboulas C, Swedani A, Ward C, Al-Madhoun AS, Wilton S, et al. HDAC activity regulates entry of mesoderm cells into the cardiac muscle lineage. J Cell Sci. 2006;119:4305–14.

    Article  Google Scholar 

  19. Takahashi-Fujigasaki J, Fujigasaki H. Histone deacetylase (HDAC) 4 involvement in both Lewy and Marinesco bodies. Neuropathol Appl Neurobiol. 2006;32:562–6.

    Article  Google Scholar 

  20. Santo L, Hideshima T, Kung AL, Tseng JC, Tamang D, et al. Preclinical activity, pharmacodynamic, and pharmacokinetic properties of a selective HDAC6 inhibitor, ACY-1215, in combination with bortezomib in multiple myeloma. Blood. 2012;119:2579–89.

    Article  Google Scholar 

  21. Bobrowska A, Paganetti P, Matthias P, Bates GP. Hdac6 knock-out increases tubulin acetylation but does not modify disease progression in the R6/2 mouse model of Huntington's disease. PLoS One. 2011;6:e20696.

    Article  Google Scholar 

  22. Destaing O, Saltel F, Gilquin B, Chabadel A, Khochbin S, et al. A novel rho-mDia2-HDAC6 pathway controls podosome patterning through microtubule acetylation in osteoclasts. J Cell Sci. 2005;118:2901–11.

    Article  Google Scholar 

  23. Gronthos S, Mankani M, Brahim J, Robey PG, Shi S. Postnatal human dental pulp stem cells (DPSCs) in vitro and in vivo. Proc Natl Acad Sci U S A. 2000;97:13625–30.

    Article  Google Scholar 

  24. Sun SZ, Liu SH, Wei FC, Zhang CY, Liu YS. The ALP activity of cultured dental coronal and root pulp cells in vitro. Shanghai Kou Qiang Yi Xue. 2004;13:531.

    PubMed  Google Scholar 

  25. Lindunger A, MacKay CA, Ek-Rylander B, Andersson G, Marks SC Jr. Histochemistry and biochemistry of tartrate-resistant acid phosphatase (TRAP) and tartrate-resistant acid adenosine triphosphatase (TrATPase) in bone, bone marrow and spleen: implications for osteoclast ontogeny. Bone Miner. 1990;10:109–19.

    Article  Google Scholar 

  26. Kusano K, Miyaura C, Inada M, Tamura T, Ito A, et al. Regulation of matrix metalloproteinases (MMP-2, −3, −9, and −13) by interleukin-1 and interleukin-6 in mouse calvaria: association of MMP induction with bone resorption. Endocrinology. 1998;139:1338–45.

    Article  Google Scholar 

  27. Choi J, Choi SY, Lee SY, Lee JY, Kim HS, et al. Caffeine enhances osteoclast differentiation and maturation through p38 MAP kinase/Mitf and DC- STAMP/CtsK and TRAP pathway. Cell Signal. 2013;25:1222–7.

    Article  Google Scholar 

  28. Ravindran S, Huang CC, George A. Extracellular matrix of dental pulp stem cells: applications in pulp tissue engineering using somatic MSCs. Front Physiol. 2014;4:395.

    Article  Google Scholar 

  29. Jewett A, Arasteh A, Tseng HC, Behel A, Arasteh H, et al. Strategies torescue mesenchymal stem cells (MSCs) and dental pulp stem cells (DPSCs) from NK cell mediated cytotoxicity. PLoS One. 2010;5:e9874.

    Article  Google Scholar 

  30. Rodas-Junco BA, Canul-Chan M, Rojas-Herrera RA, De-la-Pena C, Nic-Can GI. Stem cells from dental pulp: what epigenetics can do with your tooth. Front Physiol. 2017;8:999.

    Article  Google Scholar 

  31. Sanen K, Martens W, Georgiou M, Ameloot M, Lambrichts I, et al. Engineered neural tissue with Schwann cell differentiated human dental pulp stem cells: potential for peripheral nerve repair? J Tissue Eng Regen Med. 2017;11:3362–72.

    Article  Google Scholar 

  32. Soares DG, Rosseto HL, Scheffel DS, Basso FG, Huck C, et al. Odontogenic differentiation potential of human dental pulp cells cultured on a calcium-aluminate enriched chitosan-collagen scaffold. Clin Oral Investig. 2017;21:2827–39.

    Article  Google Scholar 

  33. Song JS, Takimoto K, Jeon M, Vadakekalam J, Ruparel NB, et al. Decellularized human dental pulp as a scaffold for regenerative endodontics. J Dent Res. 2017;96:640–6.

    Article  Google Scholar 

  34. Song M, Lee JH, Bae J, Bu Y, Kim EC. Human dental pulp stem cells are more effective than human bone marrow-derived mesenchymal stem cells in cerebral ischemic injury. Cell Transplant. 2017;26:1001–16.

    Article  Google Scholar 

  35. Gordeuk VR. Osteoclast activation and sickle bone disease. Blood. 2015;126:2259–60.

    Article  Google Scholar 

  36. Jiang H, Wang Y, Viniegra A, Sima C, McCulloch CA, et al. Adseverin plays a role in osteoclast differentiation and periodontal disease-mediated bone loss. FASEB J. 2015;29:2281–91.

    Article  Google Scholar 

  37. Mucci JM, Suqueli Garcia F, de Francesco PN, Ceci R, Di Genaro S, et al. Uncoupling of osteoblast-osteoclast regulation in a chemical murine model of Gaucher disease. Gene. 2013;532:186–91.

    Article  Google Scholar 

  38. Boyce BF, Rosenberg E, de Papp AE, Duong LT. The osteoclast, bone remodelling and treatment of metabolic bone disease. Eur J Clin Investig. 2012;42:1332–41.

    Article  Google Scholar 

  39. Singer FR, Leach RJ. Bone: do all Paget disease risk genes incriminate the osteoclast? Nat Rev Rheumatol. 2010;6:502–3.

    Article  Google Scholar 

  40. Iwami K, Moriyama T. Effects of short chain fatty acid, sodium butyrate, on osteoblastic cells and osteoclastic cells. Int J BioChemiPhysics. 1993;25:1631–5.

    Article  Google Scholar 

  41. El-Serafi AT, Oreffo RO, Roach HI. Epigenetic modifiers influence lineage commitment of human bone marrow stromal cells: differential effects of 5-aza-deoxycytidine and trichostatin a. Differentiation. 2011;81:35–41.

    Article  Google Scholar 

  42. Kim HN, Ha H, Lee JH, Jung K, Yang D, et al. Trichostatin a inhibits osteoclastogenesis and bone resorption by suppressing the induction of c-Fos by RANKL. Eur J Pharmacol. 2009;623:22–9.

    Article  Google Scholar 

  43. Ali SA, Dobson JR, Lian JB, Stein JL, van Wijnen AJ, et al. A RUNX2-HDAC1 co-repressor complex regulates rRNA gene expression by modulating UBF acetylation. J Cell Sci. 2012;125:2732–9.

    Article  Google Scholar 

  44. Schroeder TM, Kahler RA, Li X, Westendorf JJ. Histone deacetylase 3 interacts with runx2 to repress the osteocalcin promoter and regulate osteoblast differentiation. J Biol Chem. 2004;279:41998–2007.

    Article  Google Scholar 

  45. Westendorf JJ, Zaidi SK, Cascino JE, Kahler R, van Wijnen AJ, et al. Runx2 (Cbfa1, AML-3) interacts with histone deacetylase 6 and represses the p21(CIP1/WAF1) promoter. Mol Cell Biol. 2002;22:7982–92.

    Article  Google Scholar 

  46. Jensen ED, Schroeder TM, Bailey J, Gopalakrishnan R, Westendorf JJ. Histone deacetylase 7 associates with Runx2 and represses its activity during osteoblast maturation in a deacetylation-independent manner. J Bone Miner Res. 2008;23:361–72.

    Article  Google Scholar 

  47. Pham L, Kaiser B, Romsa A, Schwarz T, Gopalakrishnan R, et al. HDAC3 and HDAC7 have opposite effects on osteoclast differentiation. J Biol Chem. 2011;286:12056–65.

    Article  Google Scholar 

  48. Jin Z, Wei W, Dechow PC, Wan Y. HDAC7 inhibits osteoclastogenesis by reversing RANKL-triggered beta-catenin switch. Mol Endocrinol. 2013;27:325–35.

    Article  Google Scholar 

  49. Zhang CL, McKinsey TA, Chang S, Antos CL, Hill JA, et al. Class II histone deacetylases act as signal-responsive repressors of cardiac hypertrophy. Cell. 2002;110:479–88.

    Article  Google Scholar 

  50. Cohen-Kfir E, Artsi H, Levin A, Abramowitz E, Bajayo A, et al. Sirt1 is a regulator of bone mass and a repressor of Sost encoding for sclerostin, a bone formation inhibitor. Endocrinology. 2011;152:4514–24.

    Article  Google Scholar 

  51. Luxton GW, Gundersen GG. HDAC6-pack: cortactin acetylation joins the brew. Dev Cell. 2007;13:161–2.

    Article  Google Scholar 

  52. Zhang X, Yuan Z, Zhang Y, Yong S, Salas-Burgos A, et al. HDAC6 modulates cell motility by altering the acetylation level of cortactin. Mol Cell. 2007;27:197–213.

    Article  Google Scholar 

  53. Rao R, Fiskus W, Yang Y, Lee P, Joshi R, et al. HDAC6 inhibition enhances 17-AAG--mediated abrogation of hsp90 chaperone function in human leukemia cells. Blood. 2008;112:1886–93.

    Article  Google Scholar 

  54. Kovacs JJ, Murphy PJ, Gaillard S, Zhao X, Wu JT, et al. HDAC6 regulates Hsp90 acetylation and chaperone-dependent activation of glucocorticoid receptor. Mol Cell. 2005;18:601–7.

    Article  Google Scholar 

  55. Ota S, Zhou ZQ, Romero MP, Yang G, Hurlin PJ. HDAC6 deficiency or inhibition blocks FGFR3 accumulation and improves bone growth in a model of achondroplasia. Hum Mol Genet. 2017;26:3651.

    Article  Google Scholar 

  56. Ota S, Zhou ZQ, Romero MP, Yang G, Hurlin PJ. HDAC6 deficiency or inhibition blocks FGFR3 accumulation and improves bone growth in a model of achondroplasia. Hum Mol Genet. 2016;25:4227–43.

    Article  Google Scholar 

Download references

Acknowledgements

The authors appreciate all participants for their willingness and time to share their views and experiences, which laid the foundation for this qualitative study.

Funding

This study was supported by the Beijing Nova Program (Xxjh2015105) and the General Hospital of PLA Clinical Support Fund (2017FC-TSYS-2016).

Availability of data and materials

All data generated or analyzed during this study are included in this published article. The datasets analyzed during the current study are available from the corresponding author on reasonable request.

Author information

Authors and Affiliations

Authors

Contributions

YW contributed to the study design, data analysis, and writing the manuscript; JKC was involved in project coordination, data analysis, and manuscript preparation; ZYS contributed to the project’s conception, experimental design, and data collection; JF was involved in data interpretation, manuscript preparation, and submission of the manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Jun Kai Cao.

Ethics declarations

Ethics approval and consent to participate

Ethical approval was obtained from the PLA General Hospital Institutional Review Board (Ethics Committee). Informed written consent was obtained from all participants.

Consent for publication

Not applicable

Competing interests

The authors declare that they have no competing interests.

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, Y., Shi, Z.Y., Feng, J. et al. HDAC6 regulates dental mesenchymal stem cells and osteoclast differentiation. BMC Oral Health 18, 190 (2018). https://doi.org/10.1186/s12903-018-0624-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12903-018-0624-1

Keywords